Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

energies

Article
A Comparative Analysis of a Power System Stability with
Virtual Inertia
Lavr Vetoshkin * and Zdeněk Müller

Department of Electrical Power Engineering, Czech Technical University, 166 27 Prague, Czech Republic;
zdenek.muller@fel.cvut.cz
* Correspondence: vetoslav@fel.cvut.cz

Abstract: The paper investigates the stability of a power system with synchronverters. A synchron-
verter is a control strategy for voltage source converters that introduces virtual inertia by mimicking
synchronous machines. The authors picked a commonly known IEEE 9 bus and IEEE 39 bus test
case systems for the test case studies. The paper presents the power system’s modal analysis with
Voltage Source Converters (VSCs) controlled as synchronverters, vector control, or Rate of Change
of Frequency-based Virtual Synchronous Generator, thus comparing different approaches to VSC
control. The first case study compares selected control algorithms, the IEEE 9 bus system, with
one VSC in the paper. The results demonstrate the benefits of synchronverters over other control
strategies. The system with synchronverters has a higher minimal damping ratio, which is proven to
be the case by numerical simulations. In the second case study, the effects of virtual inertia place-
ment were investigated. The computations showed that placement is indeed important, however,
the control strategy is as important. Besides, the system with synchronverters exhibits better stability
characteristics. The paper demonstrates that the application of synchronverters is feasible and can
 meet the demand for algorithms that bring the benefits of virtual inertia.


Citation: Vetoshkin, L.; Müller, Z. Keywords: power system stability; virtual inertia; synchronverter; small-signal stability; power sys-
A Comparative Analysis of a Power tem modeling; renewables
System Stability with Virtual Inertia.
Energies 2021, 14, 3277. http://
doi.org/10.3390/en14113277
1. Introduction
Academic Editors: Akhtar Kalam and
Power systems have dramatically changed in the last decade. The penetration of
Ilan Aharon
Renewable Energy Sources (RESs) combined with plans to phase out nuclear power plants
in many countries shapes the industry. Those developments brought new dilemmas to
Received: 6 April 2021
Accepted: 1 June 2021
power system operators and planners. A major one is how to deal with low grid inertia.
Published: 3 June 2021
There is no doubt that inertia plays a significant role in short-term frequency control. Since
it naturally stores energy, which is used during a spike in a power imbalance in the gird.
Publisher’s Note: MDPI stays neutral
The importance of inertia for frequency regulation is demonstrated in Figure 1.
with regard to jurisdictional claims in
Therefore, many people in the industry are trying to solve the upcoming problem with
published maps and institutional affil- control of a low inertia grid. Some researchers proposed to change the requirements on
iations. Rate of Change of Frequency (ROCOF) as a simple solution. This approach was adopted
in Ireland [1]. There are other proposals, one that stands out the most is an introduction
of virtual inertia. For instance, Hydro-Québec started to worry about low inertia in 2003
and set up requirements for power plant operators [2]. Similar ideas were proposed across
Copyright: © 2021 by the authors.
the Atlantic by ENTSO-E in 2017 [3]. However, this document does not set mandatory
Licensee MDPI, Basel, Switzerland.
requirements but instead introduces National Power System operators’ guidelines in
This article is an open access article
Europe. Furthermore, it outlines some examples of Canada that already require frequency
distributed under the terms and response by wind power plants. Reference [2] probably preceded others in setting up
conditions of the Creative Commons virtual inertia guidelines.
Attribution (CC BY) license (https:// The most cost-effective solution to a low inertia grid is to make RES power plants
creativecommons.org/licenses/by/ behave like synchronous generators. The fundamental idea is to control VSC so it will
4.0/). respond to a disturbance like a synchronous generator. Intensive research of that topic

Energies 2021, 14, 3277. https://doi.org/10.3390/en14113277 https://www.mdpi.com/journal/energies


Energies 2021, 14, 3277 2 of 20

yielded many different control strategies. The authors primarily divide them into two
categories grid following and grid forming. The grid-following control for synchronization
requires a stiff grid reference. Vector control is typical grid-following control that requires
Phase Locked Loop (PLL) for synchronization with the grid [5]. Different proposed control
strategies emulate inertia response, however, some still do it through PLL-provided mea-
surement of frequency [6,7]. The authors call such control strategies based on frequency
measurement Rate of Change of Frequency-based Virtual Synchronous Generators or in ab-
breviation ROCOF VSG. Some of such proposed topologies [8] utilize simple power-based
calculations. Others build on vector control and introduce voltage reference signal that
emulate inertia response [9,10]. The other control approach is to emulate synchronous gen-
erators and make a VSC essentially a grid-forming device. Synchronous generator based
topologies are synchronverters [11,12] and Virtual Synchronous Machine (VISMA) [13,14].
The advantage that emulation of inertia brings that it becomes a tunable parameter in the
system. The rotating mass of the rotor gives the inertia of the generator. The energy storage
limits virtual inertia on the other hand that Renewable Energy Source has. Usually, it is a
capacitor or a battery.

Figure 1. Time intervals of frequency response during a contingency. Reprinted from [4].

Existing research papers mainly focused on analyzing the stability of a synchronverter


connected to simple networks without comparing them to other common control strategies.
For instance an interesting studies [15,16] on the application of synchronverters to damp
oscillations grid did not discuss other possible control algorithms that can be an alternative
to synchronverters. Additionally, in these papers, the dynamics of the DC bus is not con-
sidered. Furthermore, paper [17] investigates the interaction between grid following and
grid forming devices. However, the test case system is focused on the stability of power
converters, not on the power system dynamics with several generators. Thus, the main
contribution of the paper to the field of power system stability is the following. The re-
search focuses on the dynamics and stability of a power system with multiple generators.
The paper compares different VSC control strategies that represent the main categories
outlined by the authors. The chosen control strategies are vector control, ROCOF VSG [9]
and synchronverters [11]. The paper presents the comparative analysis of the investigated
control topologies in IEEE 9 bus and IEEE 39 bus benchmark systems. For the comparison,
the authors conducted modal analysis and numerical simulations to verify the conclu-
sions. In the first case, the authors analyze the IEEE 9 bus system’s stability with only one
Photovoltaic Power Plant. In the second case where the test case system is large, authors
investigate the effects of placement of VSCs with different control strategies. The conclu-
Energies 2021, 14, 3277 3 of 20

sion emerged from the placement analysis that the system’s performance depends not only
on the placement of virtual inertia but on the control strategy that implements it.

2. Mathematical Models of the Devices


The section provides mathematical models of the devices connected to the power
system. Furthermore, it explains some assumptions that have been made in the modeling.
Additionally, it briefly describes how the devices can be connected to the network.

2.1. Synchrnous Generator: Two-Axis Model


The two-axis model or 4th order model is used in the power system model described
in the book [18]. This model closely enough approximates the generator’s behavior even
though it neglects the dynamics of one damping winding in d and one in q axis. Further-
more, the stator transients are neglected [19]. This introduces some errors, however, the
results are more conservative. Thus it is not a problem for stability studies. For the ith
generator, the model can be stated as follows.

dEq,i 0
Td0,i = − Eq,i − ( xd,i − xd,i )id,i + E f d,i (1)
dt
dEd,i 0
Tq0,i = − Ed,i + ( xq,i − xq,i )iq,i (2)
dt

dδi
= ω0 ( ω i − ω s ) (3)
dt

dδi 0
Ji = Mmech,i − Eq,i iq,i − Ed,i id,i − ( xd,i − xd,i )id,i iq,i − TFW (4)
dt
Stator algebraic equations:
0
Eq,i = Rs,i iq,i + xd,i id,i + vq,i (5)

0
Ed,i = Rs,i id,i − xq,i iq,i + vd,i (6)

Eq,i , Ed,i transient voltage in d − q axis


E f d,i field voltage
iq,i , id,i current in d − q axis
Ji inertia constant
wi generator angular velocity
δi load angle
w0 nominal angular velocity
ws synchronous angular velocity
TFW damping torque to emulate the effect of damping windings

2.2. VSC Electric Part Model


Hereafter the circuit equations of the Voltage Source Converter are stated. Figure 2
shows the converter’s line diagram with an L-type filter. A further assumption is to neglect
DC side switching losses since they are not relevant for grid dynamics/interaction. Thus it
simplifies the dynamics of the DC side of the VSC.

md vdc di
= L f d,i − ωs L f iq,i + R f id,i + vd,i (7)
2 dt

mq vdc diq,i
= Lf + ωs L f id,i + R f iq,i + vq,i (8)
2 dt

dvdc p pv − pvsc
Cdc = (9)
dt vdc
Energies 2021, 14, 3277 4 of 20

pvsc = (md vdc id + mq vdc iq )/2 (10)


L f ,i filter inductance
R f ,i filter resistance
mq,i , md,i modulation vector in d − q
Cdc DC capacitor
vdc DC voltage
iq,i , id,i injected current in d − q axis
p pv power produced by PV panels
pvsc power injected to the grid
vd,i , vq,i voltage in PCC

ppv
Lf
vdc

Cdc

Figure 2. VSC schematic diagram.

2.3. Synchronverter Model in d − q Reference Frame


The paper [11] introduced the notion of a synchronverter, a converter that mimics a
synchronous generator. Figure 3 shows the block diagram of the synchronverter. This sec-
tion provides the mathematical model in d − q frame of a synchronverter connected to
the grid.

Ppv
DC voltage
measurement

PWM
modulator
Vdc,ref
Vabc,ref
ia,ib,ic
Synchronverter
Equations
Lf
Qref

va,vb,vc
Vpcc,ref

PCC
AC voltage
measurement

Figure 3. Synchronverter block diagram.

The main idea behind the synchronverter control strategy is that it has, to some
degree, mimic the behavior of a synchronous machine [20]. Hence, the mathematical model
Energies 2021, 14, 3277 5 of 20

of synchronverters in the d − q frame should be similar to the synchronous generator


equations. Assuming that the virtual machine has only three winding in d − q frame:

dψd
vd = Rd id + − ωψq (11)
dt

dψq
vq = Rq iq + + ωψd (12)
dt

dψ f
vf = Rf if + (13)
dt
The relations between fluxes and currents are following:

− Ldd
     
ψd 0 Lfd id
 ψq  =  0 − Lqq 0  ·  iq  (14)
ψf − Ld f 0 Lf f if

where L f d is the magnetic coupling inductance between filed circuit and d stator axis.
Furthermore, in synchronverters, the assumption is that there is only one-way magnetic
coupling i.e., from rotor t stator. Therefore, Ld f = 0 and Ldd = L f , L f being leakage
inductance in a real machine or it is filter inductance, in the case of the synchronverter.
Additionally, the derivative of the field current is assumed to be 0. With the assumption
that Rd = Rq = Rs Rs is filter resistance Equations (11)–(13) can be simplified as follows

did
v d = − R s i d + ωr L f i q − L f (15)
dt

diq
v q = − R s i q − ωr L f i d − L f + ωr ψ f (16)
dt
where ψ f = L f d i f
Clearly (15) and (16) resemble the equations of the VSC electrical part (7) and (8).
Thus modulations signals become md = 0 and

m q = ωr ψ f (17)

The swing equation emulating the mechanical part:

dωr p psynch
J = mech − − D p ( ωr − ω n ) (18)
dt ωr ωr

where power injected by synchronverter psynch in a synchronverter rotating frame is calcu-


lated considering assumptions (10) and (17). Thus,

ωr ψ f vdc r
psynch = iq (19)
2
Virtual mechanical power input—pmech is set to control the DC bus voltage [21,22].
Thus, the equations for the controller are:

pmech = K p,p (vdc − vdc,re f ) + preg,pi (20)

dpreg,pi
= Ki,p (vdc − vdc,re f ) (21)
dt
where K p,p and Ki,p that are proportional and integral constants respectively
Energies 2021, 14, 3277 6 of 20

The virtual load angle of the synchronverter is computed similarly to a synchronous machine:


= ωr − ω s (22)
dt
As Figure 3 shows that in PCC the synchronverter can control reactive power injection
and/or voltage. Assuming that the synchronverter is set to be P-V node then Ki,q = 0


= Ki,v (vre f − v pcc ) + Ki,q (qre f − q) (23)
dt
where v pcc is the voltage amplitude in PCC and Ki,v is integral constant of voltage regulator.

2.4. Vector Control of VSC


Voltage Source Converters are usually equipped with vector control. Vector control
block diagram is shown in Figure 4. This control strategy allows decoupling the control
of active and reactive power injections using Park transformation (A1). In power system
transient stability studies, the effects of some parts are neglected, for example, losses in
the converter. The mathematical model is provided here in addition to the equations of
the converter. Here the calculation of modulation of voltage vector is provided. It should
be noted that the control objective in the authors’ models is the voltage in PCC, not the
reactive power injection. Furthermore, the chosen model utilizes a PI regulator for DC
voltage control and an Integral (I) regulator for PCC voltage. Additionally, for the voltage
shift’s compensation due to filter resistance and reactance, the feedforward shift is used
for modulation [23]. Synchronization with the grid is provided by synchronous reference
frame PLL (SRF-PLL). A thorough derivation of equations is not necessary in this case
since it is a commonly known control strategy.

Ppv
vdc vdc
DC voltage
measurement

DC voltage id,ref current Vdq,ref dq Vabc,ref PWM pulses


Vdc,ref regulator controller abc modulator

Vpcc,ref PCC voltage iq,ref idq,mes dq ia,ib,ic


regulator
abc
Lf

vpcc AC voltage θpll va,vb,vc


PLL
measurement
PCC

Figure 4. Vector control block diagram.

md = vd − ωs L f iq + K p,id (id,re f − id ) + Md (24)


mq = vq + ωs L f iq + K p,iq (iq,re f − iq ) + Mq (25)
dMd
= Ki,id (id,re f − id ) (26)
dt
dMq
= Ki,iq (iq,re f − iq ) (27)
dt
dpreg,pi
= Ki,dc (vdc − vdc,re f ) (28)
dt
Energies 2021, 14, 3277 7 of 20

did,re f
= Ki,p ( preg,pi + K p,dc (vdc − vdc,re f ) − pvsc ) (29)
dt
diq,re f
= Ki,v (vre f − v pcc ) (30)
dt
dI pll re f
= k i,pll (vcq − vq ) (31)
dt
dω pll re f
= I pll + k p,pll (vcq − vq ) (32)
dt
dθ pll
= ω pll − ωs (33)
dt
Note, currents regulator constants are the same, thus K p,id = K p,iq and Ki,id = Ki,iq .

2.5. ROCOF VSG


The third examined strategy is somewhat between vector control and synchronverters.
It also provides an inertia response to disturbance. However, the inertia control response is
built on top of vector control. There are different topologies that emulate inertia response
by using frequency measurement. Authors name such control approach a Rate of Change of
Frequency Virtual Synchronous Generator or in abbreviation ROCOF VSG. This particular
topology was proposed in the paper [9]. Figure 5 shows the block diagram of the examined
control strategy. Fundamentally the mathematical model of this topology is similar to
vector control with only one difference. The DC bus voltage reference consists of two
signals. The first is the voltage bus reference, and the second is the virtual inertia response
signal that modifies the DC bus reference. Thus to build mathematical model of a VSC
control as ROCOF VSG in the Equations (25)–(33) DC bus voltage reference signal is going
to be:
vdc,re f = vdc,nominal + VJ,re f + D p (ω pll − ωn ) (34)
The control signal of virtual inertia response emulation is as follows:

dVJ,re f
= ( H p ω̇ pll − VJ,re f )/Tj (35)
dt
where H p emulate inertia and D p damping constants. The time delay Tj simulates the
gradual power injection from the kinetic energy of a synchronous generator. Thus, ROCOF
VSG provides a power response based on frequency change. It should be noted that this
control strategy is not grid forming in itself. It does support frequency but still uses PLL to
track the frequency of the grid. Therefore, ROCOF VSG is very sensitive to the ability of
PLL to track frequency correctly without oscillations.
Ppv
vdc vdc
DC voltage
measurement

id,ref Vdq,ref dq Vabc,ref pulses


DC voltage current PWM
Vdc,ref regulator controller abc modulator

Vpcc,ref PCC voltage iq,ref idq,mes dq ia,ib,ic


regulator
abc
Lf

θpll
va,vb,vc
PLL
PCC
vpcc AC voltage
measurement

Figure 5. ROCOF VSG (Rate of Change of Frequency Virtual Synchrnous Generator) block diagram.
Energies 2021, 14, 3277 8 of 20

2.6. Network Model


In transient stability studies, conventionally, the dynamics of lines are neglected.
Since in comparison to time constants of machines, the propagation in lines happens almost
instantly. Therefore the network equations become algebraic equations. The general form
of a power system mathematical model has the following form:

dx
= f ( x, y, t) (36)
dt
0 = g( x, y, t) (37)
where x is the state vector, which depending on model, may contain variables of current
injections, excitation controllers, turbine governors etc. Additionally, y vector usually
corresponds to voltages.
The general form of network algebraic equations that describe power flow is following

Y · V − I ( x, V ) = 0 (38)

where I ( x, V ) respects variable impedance loads. However, for simplifications often the
constant impedance loads are used in stability studies [18,24,25]. Thus, equations can
be simplified
Y · V = I (x) (39)
Figure 6 shows the relation between different reference frames. In the authors’ implemen-
tation, all devices are modeled in their own rotating frame for more modular code, and the
current injections are transformed using Park transformation Equation (A1) (Appendix A).
The transformation between different d − q rotating frames can be made directly by sim-
plifying transformation matrices. Let’s examine ith generator with angular velocity wr,i .
The network has synchronous velocity ωs , sometimes it is calculated as a fictitious center
of inertiaR reference or fixed to one of the generators. Consequently, machine electrical angle
is θr = ωr dt or θr = ωs t + δ. As Figure 6 shows to transform a vector from machine
d − q reference frame to two network d − q reference frame. Firstly, the Inverse of the
park transformation Equation (A1) where angles are machine electrical angle must be
applied. Then from abc natural frame back to network d − q frame, the park transforma-
tion Equation (A1) with grid electrical angle is applied. It can be written more strictly in
matrix form:
~f grid = Tdq,grid · T −1 ~f rotor (40)
dq,i

Assuming that zero component can be neglected, i.e., the symmetrical network:
" grid #  " rotor #
f d,i f

cos(δi ) − sin(δi )
grid = · d,i
rotor (41)
f q,i sin(δi ) cos(δi ) f q,i

Tdq,grid·Tdq,i-1
Machine dq Network dq
reference frame reference frame

Tdq,i-1
Tdq,grid

Natural abc
frame

Figure 6. Conversions between reference frames.


Energies 2021, 14, 3277 9 of 20

3. Stability Comparison
In this section, the authors compare a power system’s stability with a photovoltaic
power plant controlled as a synchronverter, or vector control strategy, or ROCOF VSG.
Additionally, a brief description of the methodology that was used for comparison is
provided. For the case studies, the authors picked a commonly know IEEE 9 bus and IEEE
39 bus test systems [26]. IEEE 9 bus is complex enough but still can be easily understood.
Therefore, for the IEEE 9 bus test system, the authors chose to carry out modal analysis
and numerical simulations. The modal analysis provides an insight into system dynamics,
and transient stability verifies whether the system can remain stable after big disturbances
such as faults. For the case study of VSCs in the IEEE 39 bus test system, authors using
modal analysis compare the quality of control for different configurations of the network.
This test system allows demonstration of the effects of placement of virtual inertia since it
is much larger. The mathematical models of the devices were implemented in Wolfram
Mathematica [27]. Wolfram Mathematica combines powerful symbolic capability with all
possible numerical tools for solving algebraic differential equations.

3.1. Modal Analysis


For small-signal stability analysis the general form of power system mathematical
model (36) and (37) is linearized. Hence, the linear form of the model is as follows:

∆ ẋ = A · ∆x + B · ∆v (42)

where ∆x is the state vector of the power system, A is the system matrix corresponding to
devices constants and B is the input matrix that correlates with node voltages. Algebraic
equations describe the power flow in the network.

∆i = YN · ∆v (43)

where YN network admittance matrix and ∆i is the vector of current injections


The current injections of the devices are the following:

∆i = CD · ∆x + DD · ∆v (44)

where CD and DD is matrix corresponding to individual devices. Thus, algebraic equations


can be simplified:
YN · ∆v = CD · ∆x + DD · ∆v (45)
Then the solution for the vector of voltages is following

∆v = (YN − DD )−1 CD · ∆x (46)

Hence, the system of differential-algebraic equations become the system of only


differential equations:

∆ ẋ = A · ∆x + B(YN − DD )−1 CD · ∆x (47)

Consequently, Asys = A + B(YN − DD )−1 CD is new system matrix with reduced alge-
braic equation. Thus, the small-signal stability can be analyzed by finding the eigenvalues
of Asys . Furthermore, the damping ratio ζ i of individual eigenvalues might be evaluated
using the following formula:
σ
ζi = q i (48)
σi + ωi2
2

where the assumed form of the ith eigenvalue is λi = σi + jωi . The eigenvalues of the
system matrix obtained by previous computations are then used for comparing the stability
of examined strategies. The Lyapunov indirect method allows gaining insights into the
system dynamics without directly solving differential equations. The ability of a linear
Energies 2021, 14, 3277 10 of 20

system to return to the steady-state after disturbance is defined by max ( Re(λ)). Further-
more, the damping ratio of eigenvalues predicts whether the oscillations provided by the
eigenvalue persist only for a short time after disturbance. Fundamentally, it is a relation
between exponential decay of the amplitude and the oscillations given by the complex
part of the eigenvalue. Moreover, as part of the modal analysis, the authors calculated the
participation factors for individual modes. A more thorough explanation can be found in
Sauer’s book [18], or Kundur’s book [24]. The participation matrix is calculated as follows:

|wk,i ||vk,i |
pk,i = (49)
∑nk=1
|wk,i ||vk,i |

where wk,i and vk,i are components of left and right eigenvectors corresponding to ith
eigenvalue. Inherently the sum of all participation factors for individual modes is equal
to unity.

3.2. Case Study: IEEE 9 Bus


Western Electricity Coordinating Council three-machine system (Figure 7) is a classic
test case system for stability studies. The model includes three machines, nine buses, and
three loads. The authors chose to use constant impedance loads in the models. Furthermore,
the machines are equipped with IEEE type 1 exciter and Steam turbine governor model
(can be found in Sauer’s book [18]). Parameters of the system can be found in Appendix B.

Figure 7. WECC three-machine, nine bus system, Reprinted from [18].

In this test case, the PS with only one VSC connected to node 6 was analyzed. The
PV power production was 0.6 pu. Firstly, the power system mathematical model was
put together using equations in Section 2. After that, the equations were linearized, and
the reduced system matrix was obtained. Computed eigenvalues of the system with syn-
chronverters, vector control and ROCOF VSG are shown in Tables 1–3 correspondingly.
Please note, only variables with a participation factor greater than 0.2 are displayed. Fur-
thermore, the participation factor matrices are visualized in Figures A1–A3 (Appendix D).
The eigenvalues with the smallest damping ratios ξ are highlighted in bold font in the
tables. For the system with synchronverters, the damping ratio ξ min = 0.0298, which was
Energies 2021, 14, 3277 11 of 20

the best result of all examined strategies. Vector control had the lowest ξ min = 0.0062
damping ratio. The application of ROCOF VSG improved it, resulting in ξ min = 0.0112.
Figure 8 visualizes some of the results for demonstration of eigenvalues placements. In the
case of synchronverters, most eigenvalues were closer to the x-axis, which means that the
damping of these frequencies was higher than in the case of vector control or ROCOF VSG.
Furthermore, the dashed lines (black for synchronverters, red for ROCOF VSG, and blue for
vector control) show the lines corresponding to the smallest damping ratio. Additionally,
in the case of synchronverters, the max ( Re(λ)) was greater than for other control strategies.
The maximum real part of the eigenvalues is a good indicator of the quality of control that
provides information about the system’s ability to return to a steady-state after disturbance.
For the ROCOF VCG and vector control, the eigenvalue with max ( Re(λ)) was introduced
by PLL.

Table 1. The eigenvalues and participation factors of the system with synchronverters.

Eigenvalues Variable PF
λ1,2 = −72.53 ± 1265.85j iq,s , Vdc 0.5, 0.2
λ3,4 = −170.41 ± 305.73j δs , id,s , ωs 0.32, 0.27, 0.22
λ5,6 = −0.64 ± 12.86j ω3 , δ3 0.38, 0.39
λ7,8 = −5.5 ± 7.95j E f d,2 , Vr,2 0.4, 0.4
λ9,10 = −5.34 ± 7.93j E f d,3 , Vr,3 0.38, 0.39
λ11,12 = −5.25 ± 7.86j E f d,1 , Vr,1 0.39, 0.41
λ13,14 = −0.25 ± 8.56j ω2 , δ2 0.29, 0.37
λ15,16 = −1.15 ± 6.69j preg,pi , Vdc 0.37, 0.32
λ17 = −5.76 Psv,2 , Psv,3 0.27, 0.38
λ18,19 = −5.26 ± 0.45j Ed,2 , Ed,3 , Psv,3 0.2, 0.24, 0.25
λ20 = −5.17 Psv,1 0.79
λ21 = −3.82 Ed,2 , Ed,2 0.39, 0.44
λ22 = −3.23 Ed,1 1
λ23 = −2.37 Mmech,1 , Mmech,2 0.44, 0.45
λ24 = −2.32 Mmech,3 0.56
λ25,26 = −1.12 ± 0.82j ω1 0.23
λ27,28 = −0.39 ± 0.97j Eq,1 0.2
λ29,30 = −0.44 ± 0.75j Eq,1 , R f ,1 0.25, 0.2
λ31,32 = −0.43 ± 0.49j Eq,3 , R f ,3 0.26, 0.31
λ33 = −0.09 ψs 0.99

15
Synchronverter
Vector Control
10
ROCOF VSG

5
Im(λ)

-5

-10

-15
-2.0 -1.5 -1.0 -0.5 0.0
Re(λ)

Figure 8. Eigenvalues of the IEEE 9 bus system with VSC connected to Node 6. Black dashed line shows
ξ min = 0.0298 of the system with synchronverters. Red dashed line corresponds to ξ min = 0.0112 for
ROCOF VSG control. Blue dashed line corresponds to ξ min = 0.0062 for vector control.
Energies 2021, 14, 3277 12 of 20

Table 2. The eigenvalues and participation factors of the system with vector control.

Eigenvalues Variable PF
λ1,2 = −34.72 ± 738.89j id,vsc , Vdc 0.49, 0.5
λ3 = −74.59 iq,vsc 0.92
λ4,5 = −0.17 ± 27.3j ω pll , θ pll 0.49, 0.49
λ6,7 = −0.61 ± 12.84j ω3 , δ3 0.38, 0.39
λ8,9 = −5.5 ± 7.95j E f d,2 , Vr,2 0.39, 0.4
λ10,11 = −5.33 ± 7.92j E f d,3 , Vr,3 0.37, 0.38
λ12,13 = −5.24 ± 7.84j E f d,1 , Vr,1 0.38, 0.39
λ14,15 = −0.12 ± 8.48j ω2 , δ2 0.31, 0.4
λ16 = −5.76 Psv,2 , Psv,3 0.26, 0.36
λ17,18 = −5.26 ± 0.45j Ed,2 , Ed,3 , Psv,3 0.21, 0.23, 0.25
λ19 = −5.16 Psv,1 0.76
λ20 = −4.28 Mq 0.53
λ21 = −3.23 Ed,1 1
λ22 = −3.04 Ed,2 , Ed,3 , Mq 0.25, 0.28, 0.3
λ23 = −2.37 Mmech,1 , Mmech,2 0.41, 0.46
λ24 = −2.31 Mmech,3 0.59
λ25,26 = −1.03 ± 0.95j ω1 0.29
λ27,28 = −0.42 ± 1.15j Eq,1 0.2
λ29,30 = −0.15 ± 1.17j id,re f , Md 0.49, 0.48
λ31,32 = −0.44 ± 0.75j Eq,1 , R f ,1 0.21, 0.28
λ33,34 = −0.43 ± 0.49j Eq,3 , R f ,3 0.24, 0.29
λ35 = −0.07 iq,re f 0.97
λ36 = −0.02 preg,pi 1
λ37 = −0.02 I pll 1

Table 3. The eigenvalues and participation factors of the system with ROCOF VSG.

Eigenvalues Variable PF
λ1,2 = −8.27 ± 739.62j id,vsc , Vdc 0.49, 0.5
λ3,4 = −1.25 ± 26.25j ω pll , θ pll 0.44, 0.45
λ5,6 = −11.41 ± 13.03j iq,vsc , Mq 0.44, 0.4
λ7,8 = −0.67 ± 12.91j ω3 , δ3 0.39, 0.4
λ9,10 = −5.49 ± 7.95j E f d,2 , Vr,2 0.38, 0.38
λ11,12 = −5.32 ± 7.91j E f d,3 , Vr,3 0.32, 0.33
λ13,14 = −5.24 ± 7.82j E f d,1 , Vr,1 0.32, 0.33
λ15,16 = −0.31 ± 8.56j ω2 , δ2 0.31, 0.39
λ17 = −5.85j Psv,2 , Psv,3 0.3, 0.32
λ18,19 = −5.26 ± 0.44j Ed,2 , Ed,3 , Psv,3 0.21, 0.22, 0.23
λ20 = −5.17 Psv,1 0.81
λ21 = −4.99 VJ,re f 0.98
λ22 = −3.29 Ed,2 , Ed,3 0.37, 0.39
λ23 = −3.23 Ed,1 1
λ24 = −2.37 Mmech,1 , Mmech,2 0.4, 0.48
λ25 = −2.31 Mmech,3 0.6
λ26,27 = −1.01 ± 0.96j ω1 0.28
λ28,29 = −0.44 ± 1.17j Eq,1 0.2
λ30,31 = −0.02 ± 1.17j id,re f , Md 0.45, 0.45
λ32,33 = −0.44 ± 0.75j Eq,1 , R f ,1 0.22, 0.28
λ34,35 = −0.43 ± 0.49j Eq,3 , R f ,3 0.24, 0.29
λ36 = −0.07 iq,re f 0.97
λ37 = −0.02 preg,pi 1
λ38 = −0.02 I pll 1
Energies 2021, 14, 3277 13 of 20

The most severe disturbance that occurs in PS is a three-phase fault. Therefore,


the authors ran a simulation with the following scenario. The three-phase fault occurred
on line 8–9. After the fault was cleared, the faulted line was disconnected. Clearing
times tcl = 0.1 s and tcl = 0.2 s were considered. The simulation results for the load
angle of the third machine are presented in Figures 9 and 10. For the tcl = 0.1 s, as
Figure 9 shows vector control and ROCOF VSG had higher load angle deviation from
steady-state value than synchronverters. Besides, the ROCOF VSG expectedly was slightly
better than vector control. For tcl = 0.2 s, the system with vector control was unstable.
ROCOF VSG allowed higher load angle deviation and had prolonged oscillations than
synchronverters. It was expected because as Table 3 shows, ROCOF VSG had lower
damping and greater max ( Re(λ)). Therefore, the overall dynamic stability of the power
system was increased by applying synchronverter control. The synchronverter increased
the system’s total inertia, thus generators sped up longer due to higher system inertia,
which resulted in lower load angle deviation from steady-state value. Since vector control
did not provide additional inertia the generator loses synchronism for short-circuit with
tcl = 0.2 s. ROCOF VSG improved the system’s performance compared to vector control,
but the resulting dynamics of the system had higher load angle deviations and oscillations
than the system with synchronverters.

100

80

60
δ(°)

40

20 Synchronverter
Vector Control
0 ROCOF VSG

0 2 4 6 8 10

t(sec)

Figure 9. IEEE 9 bus: Load angle of 3 generator during 3 phase fault for tcl = 0.1 s.

200

150

100
δ(°)

50

0 Synchronverter
Vector Control
-50 ROCOF VSG

0 2 4 6 8 10

t(sec)

Figure 10. IEEE 9 bus: Load angle of 3 generator during 3 phase fault for tcl = 0.2 s.

3.3. Case Study: IEEE 39 Bus


New England 10 machine system (Figure 11) is a well-known benchmark test system.
This test case provides a fairly complex and close to the real power system model. In
Energies 2021, 14, 3277 14 of 20

the model, two VSC were considered, and the effects of control strategy on the system’s
stability were investigated. Parameters of the system can be found in Appendix C.

Figure 11. New England 10-machine, 39 bus system, Adapted from [28].

The primary focus of the case study was to analyze the effect of placement of VSCs
in different nodes. The following scenarios were considered. Two PV power plants
connected through VSCs were placed in different nodes, excluding nodes connected to
generators. Each power plant produced 0.8 pu. The following control quality criteria were
computed: as in the previous case i.e., minimal damping and max ( Re(λ). Both allowed us
to gain insights into the dynamics of the system without solving the differential equations
numerically. Figures 12–14 show that by using synchronverters in most cases the minimal
damping ratio was improved. It should be noted that in some node pairs, the results
are missing since this configurations had unstable eigenvalues, thus were excluded from
calculation. The application of ROCOF VSG also improved damping but overall to a
smaller degree than synchronverters.

0.0110

0.0108

0.0106

0.0104

0.0102

0.0100

Figure 12. IEEE 39 bus: Damping ratios for VSCs with vector control connected to different nodes.
Energies 2021, 14, 3277 15 of 20

0.0110

0.0108

0.0106

0.0104

0.0102

0.0100

Figure 13. IEEE 39 bus: Damping ratios for VSCs with ROCOF VSG control connected to differ-
ent nodes.

0.0110

0.0108

0.0106

0.0104

0.0102

0.0100

Figure 14. IEEE 39 bus: Damping ratios for synchronverters connected to different nodes.

Figures 15–17 show the results of computation of max ( Re(λ)) for the IEEE 39 bus
system with different control strategies. This quality of control index shows that the
synchronverter improved the system’s ability to return to steady-state. Furthermore,
it also shows that indeed the placement of virtual inertia was important and different
configurations of the PS could result in substantial differences in the index value. On the
contrary, for vector control, the calculated values of the max ( Re(λ)) were the same for all
considered configurations. It should be noted that for ROCOF VSG, the results were similar
to vector control. That is explained by PLL primarily dictating the results for these control
strategies. To summarize, the quality of control not only depended on the placement of
virtual inertia [29] but on the control algorithm that it implemented. The results presented
in this paper show that the quality of control could be affected by the control strategy for
virtual inertia as much as placement.
Energies 2021, 14, 3277 16 of 20

0.0150

0.0125

0.0100

0.0075

0.0050

0.0025

Figure 15. IEEE 39 bus: max ( Re(λ)) for VSCs with vector control connected to different nodes.

0.0150

0.0125

0.0100

0.0075

0.0050

0.0025

Figure 16. IEEE 39 bus: max ( Re(λ)) for VSCs with ROCOF VSG control connected to different nodes.

0.0150

0.0125

0.0100

0.0075

0.0050

0.0025

Figure 17. IEEE 39 bus: max ( Re(λ)) for synchronverters connected to different nodes.

4. Conclusions
The paper presents a thorough comparison of power system stability with different
control strategies of VSCs. In the paper, three different control strategies were compared
using modal analysis and numerical simulations. The first case study results proved that
the synchronverter provides better damping and lower max ( Re(λ)), which shows that it
can reach steady-state faster. Furthermore, numerical simulations (Figure 9) demonstrated
that Power System with synchronverters performs better during three-phase faults than
PS with vector control or ROCOF VSG. ROCOF VSG also improves transient stability
Energies 2021, 14, 3277 17 of 20

comparing to vector control and allows the system to withstand longer clearing times, but
the synchronverter does it better, as results provided in Figure 10 show. In the second case
study, the effects of placement were analyzed in IEEE 39 bus benchmark system. The results
of the analysis yielded that for vector control and VSG ROCOF, the configuration of the
network does not change max ( Re(λ)). However, for synchronverters, this parameter is
very changed with the placement of VSCs. Furthermore, the presented results that syn-
chronverters are better than other control strategies in almost every possible configuration
in terms of damping and max ( Re(λ)). The overall conclusion is that the control strategy
that provides virtual inertia is as important as the placement of VSCs.
To summarize, the presented cases demonstrate that the power system with syn-
chronverters performs better than the one with vector control or ROCOF VSG. From the
point of modal analysis, the stability is improved by the introduction of synchronverters.
Furthermore, transient stability performance is indeed enhanced by the application of a
synchronverter. In the authors’ opinion, synchronverters are a compelling control strategy
that introduces virtual inertia to the system and meets the modern power system’s require-
ments. Additionally, virtual inertia becomes a tunable parameter that gives more flexibility
to power system operators.

Author Contributions: Conceptualization, L.V. and Z.M.; methodology, L.V.; software, L.V.; valida-
tion, Z.M.; formal analysis, L.V. and Z.M.; investigation, L.V.; resources, Z.M.; writing—original draft
preparation, L.V.; writing—review and editing, L.V.; visualization, L.V.; supervision, Z.M.; project
administration, Z.M.; funding acquisition, Z.M. Both authors have read and agreed to the published
version of the manuscript.
Funding: This research was funded by Czech Technical University in Prague grant number SGS20/
165/OHK3/3T/13.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
The following abbreviations are used in this manuscript:

AC Alternating Current
DC Direct Current
VSC Voltage scource converter
VSG Virtual Synchronous Generator
PCC Point of Common Coupling
PF Participation Factor
PLL Phase Locked Loop
PS Power System
PV Photovoltaic
ROCOF Rate of Change of Frequency
SRF Synchronous Reference Frame
RES Renewable Energy Source
IEEE Institute of Electrical and Electronics Engineers

Appendix A. Park Transformation

r  cos(θ ) cos(θ − 2π cos(θ + 2π



2 3 ) 3 )
Tdq = − sin(θ ) − sin(θ − 2π
3 ) − sin(θ + 2π
3 )
 (A1)
3 1 1 1
2 2 2
Energies 2021, 14, 3277 18 of 20

Appendix B. IEEE 9 Bus: System Parameters


Machine parameters: H1 = 23.64, H2 = 6.40, H3 = 1.50, xd,1 = 0.146, xd,2 = 0.8958,
0
xd,3 = 1.3125, xd,1 0
= 0.0608, xd,2 0
= 0.1198, xd,3 = 0.1813, xq,1 = 0.0969, xq,2 = 0.8645,
0 0
xq,3 = 1.2578 , xq,1 = 0.0969, xq,2 = 0.1969, xq,3 0 = 0.25 , Tdo,1 = 8.96, Tdo,2 = 6.0,
Tdo,3 = 5.89 , Tqo,1 = 0.31, Tqo,2 = 0.535, Tqo,3 = 0.6
VSC parameters: Sn = 100 MVA, L f = 9.737 mH, R f = 147.402 µOhm, Cdc = 722 mF,
Vdc = 40 kV.
Synchronverter parameters: J = 0.2, D p = 0.15, Ki,v = 0.15, K p,p = 57.32 · 10−4 ,
Ki,p = 859.87 · 10−4
Vector Control: Ki,v = 0.64, K p,id = 38.2 · 10−4 , Ki,id = 0.1433, Ki,p = 1.43, K p,dc = 3.82,
Ki,dc = 795.78 · 10−4 , Ki,pll = 7.03619 · 10−6 , K p,pll = 63.33 · 10−4
ROCOF VSG: D p = 0.1, Tj = 0.2, H p = 5. All other parameters similar to vector control.

Appendix C. IEEE 39 Bus: System Parameters


VSC parameters: Sn = 100 MVA, L f = 9.737 mH, R f = 147.402 µOhm, Cdc = 722 mF,
Vdc = 40 kV.
Synchronverter parameters: J = 0.25, D p = 0.25, Ki,v = 0.15, K p,p = 57.32 · 10−4 ,
Ki,p = 859.87 · 10−4
Vector Control: Ki,v = 0.64, K p,id = 38.2 · 10−4 , Ki,id = 0.1433, Ki,p = 1.43, K p,dc = 3.82,
Ki,dc = 477 · 10−4 , Ki,pll = 7.03619 · 10−6 , K p,pll = 63.33 · 10−4
ROCOF VSG: D p = 0.15, Tj = 0.3, H p = 5. All other parameters similar to
vector control.

Appendix D. Figures

id,s
iq,s
Vdc 1.0
ws
ψs
Ed,1
Ed,2
Ed,3
Efd,1 0.8
Efd,2
Efd,3
Eq,1
Eq,2
Eq,3
Mmech,1 0.6
Mmech,2
Mmech,3
Preg,pi
Psv,1
Psv,2
Psv,3 0.4
Rf,1
Rf,2
Rf,3
Vr,1
Vr,2
Vr,3 0.2
w1
w2
w3
δ2
δ3
δs 0
Mode 1
Mode 2
Mode 3
Mode 4
Mode 5
Mode 6
Mode 7
Mode 8
Mode 9
Mode 10
Mode 11
Mode 12
Mode 13
Mode 14
Mode 15
Mode 16
Mode 17
Mode 18
Mode 19
Mode 20
Mode 21
Mode 22
Mode 23
Mode 24
Mode 25
Mode 26
Mode 27
Mode 28
Mode 29
Mode 30
Mode 31
Mode 32
Mode 33

Figure A1. IEEE 9 bus: Matrix of Participation factors for the system with syncrhonverter.
Energies 2021, 14, 3277 19 of 20

id,ref
Ipll
iq,ref 1.0
Md
Mq
Vdc
Ed,1
Ed,2
Ed,3
Efd,1 0.8
Efd,2
Efd,3
Eq,1
Eq,2
Eq,3
id,vsc
iq,vsc 0.6
Mmech,1
Mmech,2
Mmech,3
Preg,pi
Psv,1
Psv,2
Psv,3 0.4
Rf,1
Rf,2
Rf,3
Vr,1
Vr,2
Vr,3 0.2
w1
w2
w3
wpll
δ2
δ3
θpll 0
Mode 1
Mode 2
Mode 3
Mode 4
Mode 5
Mode 6
Mode 7
Mode 8
Mode 9
Mode 10
Mode 11
Mode 12
Mode 13
Mode 14
Mode 15
Mode 16
Mode 17
Mode 18
Mode 19
Mode 20
Mode 21
Mode 22
Mode 23
Mode 24
Mode 25
Mode 26
Mode 27
Mode 28
Mode 29
Mode 30
Mode 31
Mode 32
Mode 33
Mode 34
Mode 35
Mode 36
Mode 37
Figure A2. IEEE 9 bus: Matrix of Participation factors for the system with vector control.

id,ref
Ipll
iq,ref
Md 1.0
Mq
VJ,ref
Vdc
Ed,1
Ed,2
Ed,3 0.8
Efd,1
Efd,2
Efd,3
Eq,1
Eq,2
Eq,3
id,VSC 0.6
iq,VSC
Mmech,1
Mmech,2
Mmech,3
Preg,pi
Psv,1
Psv,2 0.4
Psv,3
Rf,1
Rf,2
Rf,3
Vr,1
Vr,2
Vr,3 0.2
w1
w2
w3
wpll
δ2
δ3
θpll 0
Mode 1
Mode 2
Mode 3
Mode 4
Mode 5
Mode 6
Mode 7
Mode 8
Mode 9
Mode 10
Mode 11
Mode 12
Mode 13
Mode 14
Mode 15
Mode 16
Mode 17
Mode 18
Mode 19
Mode 20
Mode 21
Mode 22
Mode 23
Mode 24
Mode 25
Mode 26
Mode 27
Mode 28
Mode 29
Mode 30
Mode 31
Mode 32
Mode 33
Mode 34
Mode 35
Mode 36
Mode 37
Mode 38

Figure A3. IEEE 9 bus: Matrix of Participation factors for the system with ROCOF VSG.

References
1. EirGrid/SONI. RoCoF Modification Proposal—TSOs’ Recommendations; EirGrid/SONI: Dublin, Ireland, 2012.
2. Hydro-Québec. Technical Requirements for the Connection of Generation Facilities to the Hydro-Québec Transmission System;
Hydro-Québec: Montréal, QC, Canada, 2003 .
3. ENTSO-E. Need for Synthetic Inertia (SI) for Frequency Regulation; ENTSO-E: Brussels, Belgium, 2017.
Energies 2021, 14, 3277 20 of 20

4. Kerdphol, T.; Rahman, F.S.; Watanabe, M.; Mitani, Y.; Turschner, D.; Beck, H.P. Enhanced Virtual Inertia Control Based on
Derivative Technique to Emulate Simultaneous Inertia and Damping Properties for Microgrid Frequency Regulation. IEEE Access.
2019, 7, 14422–14433 . [CrossRef]
5. Milano, F.; Dörfler, F.; Hug, G.; Hill, D.J.; Verbič, G. Foundations and Challenges of Low-Inertia Systems (Invited Paper) (Invited
Paper). In Proceedings of the 5th 2018 Power Systems Computation Conference (PSCC), Dublin, Ireland, 11–15 June 2018;
pp. 1–25. [CrossRef]
6. Sakimoto, K.; Miura, Y.; Ise, T. Stabilization of a power system with a distributed generator by a Virtual Synchronous Generator
function. In Proceedings of the 8th International Conference on Power Electronics-ECCE Asia, Jeju, Korea, 30 May–3 June 2011;
pp. 1498–1505.
7. Wu, H.; Ruan, X.; Yang, D.; Chen, X.; Zhao, W.; Lv, Z.; Zhong, Q. Small-Signal Modeling and Parameters Design for Virtual
Synchronous Generators. IEEE Trans. Ind. Electron. 2016, 63, 4292–4303. [CrossRef]
8. Karapanos, V.; de Haan, S.; Zwetsloot, K. Real time simulation of a power system with VSG hardware in the loop. In Proceedings of
the IECON 2011—37th Annual Conference of the IEEE Industrial Electronics Society, Melbourne, VIC, Australia, 7–10 November
2011; pp. 3748–3754. [CrossRef]
9. Saeedian, M.; Pournazarian, B.; Seyedalipour, S.S.; Eskandari, B.; Pouresmaeil, E. Emulating Rotational Inertia of Synchronous
Machines by a New Control Technique in Grid-Interactive Converters. Sustainability 2020, 12, 5346. [CrossRef]
10. Fang, J.; Lin, P.; Li, H.; Yang Y.; Tang, Y. An Improved Virtual Inertia Control for Three-Phase Voltage Source Converters Connected
to a Weak Grid. IEEE Trans. Power Electron. 2019, 34, 8660–8670. [CrossRef]
11. Zhong, Q.; Weiss, G. Synchronverters: Inverters That Mimic Synchronous Generators. IEEE Trans. Ind. Electron. 2011, 58, 1259–1267.
[CrossRef]
12. Zhong, Q.; Nguyen, P.; Ma, Z.; Sheng, W. Self-Synchronized Synchronverters: Inverters without a Dedicated Synchronization
Unit. IEEE Trans. Power Electron. 2014, 29, 617–630. [CrossRef]
13. Mo, O.; D’Arco, S.; Suul, J.A. Evaluation of Virtual Synchronous Machines with Dynamic or Quasi-Stationary Machine Models.
IEEE Trans. Ind. Electron. 2017, 64, 5952–5962. [CrossRef]
14. Beck, H.-P.; Hesse, R. Virtual Synchronous Machine. In Proceedings of the 9th International Conference on Electrical Power
Quality and Utilisation, Barcelona, Spain, 9–11 October 2007.
15. van Emmerik, E.L.; França, B.W.; Aredes, M. A synchronverter to damp electromechanical oscillations in the Brazilian transmission
grid. In Proceedings of the 2015 IEEE 24th International Symposium on Industrial Electronics (ISIE), Buzios, Brazil, 3–5 June 2015;
pp. 221–226. [CrossRef]
16. Brown, E.; Weiss, G. Using synchronverters for power grid stabilization. In Proceedings of the 2014 IEEE 28th Convention of
Electrical & Electronics Engineers in Israel (IEEEI), Eilat, Israel, 3–5 December 2014; pp. 1–5. [CrossRef]
17. Rosso, R.; Engelken, S.; Liserre, M. Robust Stability Investigation of the Interactions Among Grid-Forming and Grid-Following
Converters. IEEE J. Emerg. Sel. Top. Power Electron. 2020, 8, 991–1003. [CrossRef]
18. Sauer, P.W.; Pai, M.A. Power System Dynamics and Stability; Prentice Hall: Hoboken, NJ, USA, 1998.
19. IEEE Guide for Synchronous Generator Modeling Practices and Applications in Power System Stability Analyses. IEEE Std
2003, 1110–2002 . [CrossRef]
20. Zhong, Q. Four-quadrant operation of AC machines powered by inverters that mimic synchronous generators. In Proceedings of
the 5th IET International Conference on Power Electronics, Machines and Drives (PEMD 2010), Brighton, UK, 19–21 April 2010;
pp. 1–6. [CrossRef]
21. Vetoshkin, L.; Müller, Z. A comparative study of synchronverter stability. In Proceedings of the 2020 21st International Scientific
Conference on Electric Power Engineering (EPE), Prague, Czech Republic, 19–21 October 2020.
22. Vetoshkin, L.; Müller, Z. A supervisory MPC for synchronverter. In Proceedings of the 2020 21st International Scientific Conference
on Electric Power Engineering (EPE), Prague, Czech Republic, 19–21 October 2020.
23. Papangelis, L.; Debry, M.; Prevost, T.; Panciatici, P.; Van Cutsem, Y. Stability of a Voltage Source Converter Subject to Decrease
of Short-Circuit Capacity: A Case Study. In Proceedings of the 2018 Power Systems Computation Conference (PSCC), Dublin,
Ireland, 11–15 June 2018; pp. 1–7. [CrossRef]
24. Kundur, P. Power System Stability and Control; McGraw-Hill: New York, NY, USA , 1994.
25. Anderson, P.; Fouad, A. Power System Control and Stability; IEEE Press: Piscataway, NJ, USA, 2003.
26. Dembart, B.; Erisman, A.; Cate, E.; Epton, M.; Dommel, H. Power System Dynamic Analysis: Phase I. Final Report. 1977.
Available online: https://www.osti.gov/biblio/7296152 (accessed on 2 June 2021).
27. Wolfram Research, Inc. Mathematica; Version 12.2; Wolfram Research, Inc.: Champaign, IL, USA, 2020.
28. Hiskens, I. IEEE PES Task Force on Benchmark Systems for Stability Controls. 2013. Available online: http://www.sel.eesc.usp.br/ieee/
IEEE39/New_England_Reduced_Model_(39_bus_system)_MATLAB_study_report.pdf (accessed on 10 September 2020).
29. Poolla, B.K.; Bolognani, S.; Dörfler, F. Optimal Placement of Virtual Inertia in Power Grids. IEEE Trans. Autom. Control
2017, 62, 6209–6220. [CrossRef]

You might also like